Next Article in Journal
Highly Efficient Extracellular Production of Recombinant Streptomyces PMF Phospholipase D in Escherichia coli
Next Article in Special Issue
Nanocatalysts for Hydrogen Production
Previous Article in Journal
Biocatalysis at Extreme Temperatures: Enantioselective Synthesis of both Enantiomers of Mandelic Acid by Transesterification Catalyzed by a Thermophilic Lipase in Ionic Liquids at 120 °C
Previous Article in Special Issue
Light Cycle Oil Source for Hydrogen Production through Autothermal Reforming using Ruthenium doped Perovskite Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Demonstration of the Superiority of the Dual Ni-Based Catalytic System for the Adjustment of the H2/CO Ratio in Syngas for Green Fuel Technologies

by
Suntorn Sangsong
1,2,
Tanakorn Ratana
1,2,
Sabaithip Tungkamani
1,2,
Thana Sornchamni
3,
Monrudee Phongaksorn
1,2,* and
Eric Croiset
4
1
Department of Industrial Chemistry, Faculty of Applied Science, King Mongkut’s University of Technology North Bangkok, Bangkok 10800, Thailand
2
Research and Development Center for Chemical Engineering Unit Operation and Catalyst Design (RCC), King Mongkut’s University of Technology North Bangkok, Bangkok 10800, Thailand
3
Innovation Institute PTT, Sananp Thuep, Wang Noi, Phra Nakhon Si Ayutthaya 13170, Thailand
4
Department of Chemical Engineering, University of Waterloo, Waterloo, ON N2L 3G1, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(9), 1056; https://doi.org/10.3390/catal10091056
Submission received: 12 August 2020 / Revised: 8 September 2020 / Accepted: 10 September 2020 / Published: 14 September 2020
(This article belongs to the Special Issue Nanocatalysts for Hydrogen Production)

Abstract

:
A novel dual Ni-based catalytic process (DCP) to control the H2/CO ratio of 2 in the syngas product within one step at temperature <700 °C was created and constructed. With the sequence of the catalysts located in the single reactor, the endothermic combined steam and CO2 reforming of methane (CSCRM) reaction and the exothermic ultra-high-temperature water–gas shift (UHT-WGS) reaction work continuously. During the process, the H2/CO ratio is raised suddenly at UHT-WGS after the syngas is produced from CSCRM, and CSCRM utilizes the heat released from UHT-WGS. Due to these features, DCP is more compact, enhances energy efficiency, and thus decreases the capital cost compared to reformers connecting with shift reactors. To prove this propose, the DCP tests were done in a fixed-bed reactor under various conditions (temperature = 500, 550, and 600 °C; the feed mixture (CH4, CO2, H2O, and N2) with H2O/(CH4 + CO2) ratio = 0.33, 0.53, and 0.67). According to the highest CH4 conversion (around 65%) with carbon tolerance, the recommended conditions for producing syngas with the H2/CO ratio of 2 as a feedstock of Fischer–Tropsch synthesis include the temperature of 600 °C and the H2O/(CH4 + CO2) ratio of 0.53.

1. Introduction

The ongoing contamination of air by greenhouse gases has been a critical issue as it causes adverse environmental impacts around the world. GTL (gas-to-liquids) is a process that is considered as an alternative energy processing technology to convert natural gas into clean burning-liquid fuels such as gasoline, jet fuel, and diesel [1,2,3]. The one essential chemical reaction used in the commercial GTL process is the Fischer–Tropsch synthesis (FTs) which requires syngas (a mixture of H2 and CO) with the H2/CO ratio of about 2 [4,5]. Commercially, autothermal reforming and partial oxidation of natural gas, a major component of methane, are utilized to produce syngas with the H2/CO ratio of 2 in the GTL process [6,7]. However, these reforming processes have some drawbacks, such as the requirement of an oxygen unit or oxygen-enriched facilities to adjust the H2/CO ratio and the difficulty in process control due to the presence of hot spots (a highly exothermic reaction), which brings about the potential explosion danger [6,8,9,10,11]. Consequently, the production of practical syngas is necessary to overcome the drawbacks of conventional reforming processes [12]. For this approach, the combined steam and CO2 reforming of methane (CSCRM) has received considerable attention for syngas production in order to control the H2/CO ratio of syngas by adjusting the feed composition of H2O and CO2 without the additional units of the oxygen supply [13,14].
Ni-based catalysts have been used for the CSCRM process because of their reasonable prices and acceptable activity compared to the expensive noble metal catalysts (Ru, Re, Rh, Pd, Pt, Ir) [1]. The barrier to this commercial process is the rapid deactivation of Ni catalysts due to carbon deposition and metal sintering during the reaction when the S/C ratio of the feed is lower than unity [15,16]. However, the CSCRM reaction is an intensive endothermic reaction that consumes a high amount of energy for the evaporation of a large amount of water and has a (high) operating temperature typically above 700 °C to produce the syngas with H2/CO of around 2 [17,18,19]. For example, Jang et al. [20] investigated the CSCRM reaction (over Ni-MgO-Ce0.8Zr0.2O2 catalyst) with the effects of the (CO2 + H2O)/CH4 ratio (0.9–2.9), the CO2:H2O ratio (3:1–1:3), and the temperature range of 500 to 1000 °C. The authors concluded that with (CO2 + H2O)/CH4 ratios above 1.2, the CO2:H2O ratio of 1:2.1 and a temperature of at least 850 °C are preferable reaction conditions for the syngas production in the GTL process. Danilova et al. [21] studied the reaction of the CSCRM reaction under the atmospheric pressure at 750 °C over porous nickel-based catalysts. The results revealed that CH4 conversion of 94%, the syngas yield of 90%, and the higher H2/CO ratio (2.7–2.8) were achieved. Huang et al. [22] reported the effect of a steam in a feed composition on the activity of the 10wt%Ni/3wt%MgO/SBA-15 catalyst for the CSCRM reaction operated at 850 °C. The optimal CH4:CO2:H2O molar ratio was 2:1:1.5 for CH4 conversion of 98.7 %, the CO2 conversion of 92%, and the H2/CO ratio of approximately 1.79. Jabbour et al. [23] developed “one-pot” mesoporous nickel alumina-based catalysts using the EISA method for the CSCRM reaction operating at 800 °C. These high Ni dispersion catalysts enhanced activity and stability (CH4 conversion of around 80%) with a sustainable H2/CO molar ratio close to 2 in gas production. Bae et al. [24] studied the role of the CeO2–ZrO2 distribution on the Ni/MgAl2O4 catalyst in the CSCRM reaction at 850 °C. The homogeneous distribution of CeO2–ZrO2 on the Ni/MgAl2O4 catalyst demonstrated the highest CH4 conversion of 86% and the highest CO2 conversion of 58% with the H2/CO ratio of 2.2.
The high operating temperature in the CSCRM reaction may result in the cost of the synthesis gas production being approximately 60% of the total capital cost [25]. Therefore, to reduce the operating cost in the syngas production and to lessen the concern about the catalyst deactivation under severe operating conditions, energy efficiency (low operating temperature) in the syngas production process should be developed [26,27,28]. An alternative solution would be combining the CSCRM operating at a relatively low S/C ratio with a mildly exothermic process (such as the UHT-WGS and/or partial oxidation of methane) without the supply of heat in an adiabatic reactor, and thus, making the process more energy efficient [29,30]. Nevertheless, the H2/CO ratio is fixed at about 2 by adjusting the operating parameter to meet the requirements of the FTs. Furthermore, the catalyst deactivation could be suppressed by the milder operating conditions and adding an oxygen source in the feed.
In previous work [31], 5wt%Ni5wt%Co/MgO–Al2O3 (NiCo/Mg-Al) showed the higher metal dispersion, smaller metal particle size, and a high reducibility due to the effect of the metal–metal interaction. As a result, a valuable CRM catalytic activity in terms of CH4 and CO2 conversion was attained. The maintainability of the UHT-WGS catalytic performance at the temperature range of 500–600 °C over the catalyst with the composition of 10wt%Ni/5wt%CeO2-Al2O3 (Ni/5Ce-Al) was also successful because of the high surface area, high metal dispersion, and practical Ni–Ce–Al interaction [32]. Using these developed Ni-based catalysts, the novel concept of the superior dual Ni-based catalytic process (DCP) that converts CH4 and CO2, the main greenhouse gases, into syngas with the H2/CO ratio of about 2 at relatively low temperatures (<700 °C) was developed in this work. In DCP, the Ni/5Ce-Al catalyst is located next after the NiCo/Mg-Al catalyst. As a result, the CSCRM and the UHT-WGS are operated continuously in a single reactor. During the operation, the syngas is produced on the catalyst for the CSCRM and then the H2/CO ratio in the syngas is increased on the catalyst for the UHT-WGS. Moreover, the exothermic WGS can supply energy to the endothermic reactions above that take place on the CSCRM zone. Consequently, this original DCP is an alternative compact catalyst system that provides a very favorable H2/CO ratio for the syngas product associated with energy efficiency. Based on the reactants in the feed (CH4:CO2:H2O:N2) and the thermodynamic favorability, the possible reactions are demonstrated in the Equations (1)–(4).
CH 4 + CO 2 2 H 2 + 2 CO , Δ H 298   K =   + 261   kJ / mol
CH 4 + H 2 O 3 H 2 + CO , Δ H 298   K   =   + 206   kJ / mol
CH 4 + 2 H 2 O 4 H 2 + CO 2 , Δ H 298   K   =   + 165   kJ / mol
CO + H 2 O H 2 + CO 2 , Δ H 298   K   =   41   kJ / mol
Regarding Equations (1)–(4), three representative reactions are endothermic except for Equation (4). The reaction temperature and the feed composition are significant to control the activity and the composition of the syngas product.
In this work, the DCP was constructed and demonstrated. All supported catalysts were prepared by the impregnation method. The physicochemical properties of prepared catalysts were characterized. The CRM and UHT-WGS catalytic performances were separately tested with temperature programmed from 500 to 600 °C. According to the well-known stoichiometry of the CSCRM equation (3CH4 + 2H2O + CO2 8H2 + 4CO) [6,14,19], the H2/CO ratio of 2 can be achieved when managing the composition of the gas feed with S/C ratio close to 0.5 and performing under a severe temperature (≥700 °C). Then, the effects of the operating condition adjustments on the DCP catalytic performance and the H2/CO ratio were examined for various temperatures (500, 550, and 600 °C) and the feed compositions (CH4:CO2:H2O:N2 molar ratio = 1:0.5:x:1; x = 0.5, 0.8, and 1 reflecting the S/C ratios of 0.33, 0.53, and 0.67, respectively).

2. Results and Discussion

2.1. Catalyst Characterization

2.1.1. Morphological Characterization

The diffraction patterns of the calcined NiCo/Mg-Al and Ni/5Ce-Al catalysts were investigated through the XRD as displayed in Figure 1. The NiCo/Mg-Al catalyst revealed characteristic diffraction peaks of MgAl2O4 spinel at 2 θ angles of 36.8° 44.8°, and 65.3° (JCPDS 77-0435); these peaks could also be assigned to the spinel phases of NiAl2O4 (JCPDS 78-0552) and CoAl2O4 (JCPDS 82-2243) because Ni2+, Co2+, and Mg2+ can be incorporated into the identical lattice with Al2O3 [33]. However, these diffraction peaks were not easily to be identified because of either lower calcination temperatures or existing overlaps between the diffraction peaks of NiAl2O4 (or CoAl2O4) and the peaks of MgAl2O4 [34,35].
The Ni/5Ce-Al catalyst showed characteristic diffraction peaks of γ-Al2O3 at 2 θ angles of 37.5°, 45.6°, and 66.6° (JCPDS 50-0741). At the same time, the broad peaks at 2 θ angles of 28.5° and 47.5° (JCPDS 34-0394) are attributed to the cubic fluorite type structure of CeO2. Diffraction peaks that correspond to the crystalline species of NiO (JCPDS 89-7131) and Co3O4 (JCPDS 76-1802) in cubic structures were not observed for all calcined catalysts. This occurrence indicates that the active metal species transformed into the spinel structure (especially the NiCo/Mg-Al catalyst) or the high dispersion of the active metal species on the surface of the support [36].
The N2 adsorption–desorption isotherms of all catalysts are illustrated in Figure 2. According to the IUPAC classification, the NiCo/Mg-Al and Ni/5Ce-Al catalysts showed type IV isotherm curves with different hysteresis loops. The loop of NiCo/Mg-Al catalyst indicates a H3 hysteresis behavior associated with solids containing aggregates or agglomerations of particles, representing slit-like pores (plates or edged particles like cubes) with a non-uniform size and/or shape [2]. The H2 hysteresis behavior observed on Ni/5Ce-Al catalyst refers to pores with narrow mouths and an ink-bottle shape [37,38]. The pore size distributions of the NiCo/Mg-Al and the Ni/5Ce-Al catalysts comprise a mesoporous material with pore diameter ranges of 3–30 nm and 5–10 nm, respectively.
The structural properties of calcined catalysts are summarized in Table 1. The NiCo/Mg-Al catalyst has a surface area of 130 m2 g−1 and a pore volume of 0.4 cm3 g−1, and an average pore size diameter of 11.2 nm, which is in agreement with other studies [2,39]. The Ni/5Ce-Al catalyst presented the surface area of 183 m2 g−1 with a pore volume of 0.6 cm3 g−1 and an average pore size diameter of 12.5 nm, which is located in the same range compared to other published research [36,40,41]. Moreover, these two catalysts disclosed the small metal particle size due to the high metal dispersion, as explained by the XRD results.

2.1.2. Physical Characterization

Figure 3 presents the H2-TPR profiles of NiCo/Mg-Al and Ni/5Ce-Al catalysts. The TPR profile of NiCo/Mg-Al catalyst displayed shoulder peaks at lower temperatures (a range of 150 to 450 °C), which correlate to the reduction of Co3O4 to CoO and CoO to metallic Co0. The broad peak centered at 520 °C relates to the reduction of Ni2+ to Ni0 [3]. The peak at a high temperature of around 840 °C indicates the reduction of Ni or Co species with strong interactions due to the metal alloy effect and/or SMSI effect regarding the NiCo-based catalyst [33]. A similar trend of H2-TPR behavior was also reported in Li et al. [42], indicating high reduction temperatures because of the formation of NiCo alloy phases. Furthermore, the reduction peak at a high temperature (825 °C) with a NiCo catalyst has been assigned as the reduction of small active metal particles and possibly the reduction of nickel and cobalt aluminate-like compounds (NiAl2O4 and CoAl2O4 spinel structures) [43].
For the TPR profile of the Ni/5Ce-Al catalyst, small broad peaks from 200 °C to 400 °C were observed. The peak at around 270 °C can be ascribed to the reduction of the CeO2 and the NiO interacting with the partial bulk CeO2, while the peak at around 370 °C correlates to the reduction of free NiO on the catalyst support [41]. In accordance with the XRD analyses from Figure 1, the high intensity peak at 700 °C illustrates the reduction of Ni2+ ions in the amorphous spinel phases with non-stoichiometry of nickel aluminate (NiAlxOy) and stoichiometry of nickel aluminate (NiAl2O4) [44]. Actual H2-uptakes calculated from H2-TPR profiles for the prepared catalysts were close to the theoretical H2-uptake, are listed in Table 1. The result implies that the Ni2+ species were fully reduced. Although the TPR profiles suggest a reduction temperature of over 650 °C, the temperature for the reduction of all catalysts was limited at 650 °C to avoid the agglomeration of active metal according to the calcination temperature of the NiCo/Mg-Al catalyst [45].

2.2. Catalytic Performance

2.2.1. Catalytic Performance for CRM and UHT-WGS Reactions

The CRM and the UHT-WGS reactions were performed separately over NiCo/Mg-Al and Ni/5Ce-Al catalysts, respectively, for 8 h at each operating temperature (500, 550, and 600 °C). The CH4 conversion obtained from the NiCo/Mg-Al catalyst with different temperatures is provided in Figure 4a. As seen in Figure 4a, the CH4 conversion increased from ~29% at 500 °C to ~52% at 600 °C because of a highly endothermic process by nature. Furthermore, the sustainable CH4 conversion as functions of time-on-stream for each studied temperature was observed and also close to the thermodynamic equilibrium conversions (23% at 500 °C, 39% at 550 °C, and 59% at 600 °C) calculated using the reactivity test conditions. These results are in good agreement with the published literature when considering the same studied temperature range [41,46,47].
The catalytic activity of the Ni/5Ce-Al catalyst in the UHT-WGS reaction is presented in Figure 4b. It was found that the CO conversion performance decreased from ~65% at 500 °C to ~55% at 600 °C due to an exothermic reaction. Although the CO conversions obtained from UHT-WGS tests are different from the thermodynamic equilibrium conversions (92% at 500 °C, 89% at 550 °C, and 86% at 600 °C), a similar tendency was observed in the case of increasing operating temperature. Nonetheless, this catalytic behavior demonstrated the appropriate range of CO conversion and an expected trend of CO conversion compared to other studies [48,49,50]. The maintainability of CO conversion in each studied temperature was also attained. To further study the combination of the CSCRM and the UHT-WGS reactions as a dual catalytic process, the operating temperature should be limited at 600 °C to maintain the activity of the UHT-WGS due to its nature of exothermic reaction catalyst, while the CH4 conversion from the methane reforming reaction (this case refers to CRM result) is acceptable [34].

2.2.2. Catalytic Performance for DCP

Figure 5a,b exhibits the CH4 and CO2 conversions against time-on-stream with different operating temperatures (500, 550, and 600 °C) at a fixed S/C ratio of 0.67. As seen in Figure 5a, the CH4 conversion increases when the operating temperature increases because the CSCRM is the endothermic process and high temperature encourages the reactions [12]. This trend was also found for CO2 conversion (Figure 5b); the extent of the CO2 conversion rose from ~15% at 500 °C to ~35% at 600 °C. It should be noted that the appearance of the low CO2 conversion at 500 °C could involve the fact that CO2 is a product of the WGS reaction. Nevertheless, these CH4 and CO2 conversions were reliable compared to the results at a similar operating temperature range demonstrated by previously published works [2,12,51]. The obtained H2/CO ratios for different operating temperatures, displayed in Figure 5c, were close to 2 for all operating temperatures according to these conditions.
The influence of the feed composition (S/C ratio = 0.33, 0.53, and 0.67) for the DCP was investigated. Reactant conversions and the H2/CO ratio at a fixed operating temperature of 600 °C (Figure 6) illustrate that an increased S/C ratio from 0.33 to 0.53 resulted in an increase in the CH4 conversion from ~55 % to ~65 % (Figure 6a). This result implies the raising of reforming rates by the addition of the oxidizing agents [34]. Alternatively, the increased S/C ratio from 0.53 to 0.67 decreased the CH4 conversion from ~65 % to ~50 %, suggesting an excess adsorbed steam on the catalyst surface. This behavior could be explained by the work of Gensterblum et al. [52], who reported that gas sorption (CH4 or CO2) capacity decreases with increasing H2O content on coal surfaces, and the interaction of H2O with natural coal is more complex than the interactions of non-polar gases (CH4, CO2, and N2). It is speculated that the duration of the H2O dissociative adsorption is longer than said duration for other gases, which agrees with the results of Tan et al. [53]. They revealed that the adsorption energies of the H2O molecules (61–80 kJ mol−1) on metal-MOF-74 (metal = Ni, Co, and Mg) are higher than that of the adsorption energies of CH4 (19–20 kJ mol−1) and CO2 (37–48 kJ mol−1) molecules. Zhou et al. [54] also pointed out that the adsorption capacity of H2O molecules (7.552 mmol g−1) is higher than that of the adsorption capacities of CH4 (0.001 mmol g−1) and CO2 (0.241 mmol g−1) molecules, resulting in a significant decrease of the adsorption sites for the CH4 and CO2 molecules at a higher H2O content. Nevertheless, the experimental CH4 conversions for all feed compositions were slightly higher than those of the thermodynamic equilibrium (45% for S/C ratio of 0.33, 51% for S/C ratio of 0.53, and 55% for S/C ratio of 0.67).
Figure 6b represents the CO2 conversion as a function of time-on-stream. The CO2 conversion decreased from approximately 52% to 35% with an increasing S/C ratio from 0.33 to 0.67. The increase in S/C ratio caused a considerable decrease of CO2 conversion. This circumstance can be attributed to CH4 reacting with H2O instead of the CO2 at high H2O content in the feed composition (S/C ratio), since both CO2 and H2O act as co-oxidants in the DCP, but CO2 is more stable (its thermodynamic effect) with an increasing amount of H2O in the feed gas [22]. Meanwhile, the CO2 conversion in the thermodynamic equilibrium (35% for S/C ratio of 0.33, 17% for S/C ratio of 0.53, and 6% for S/C ratio of 0.67) decreased with the increase of the S/C ratio and lower than experimental CO2 conversions for all S/C ratios. The difference between the experimental and the thermodynamic equilibrium results could be caused by the side reactions (RWGS reaction) considering the similar reaction temperature [55]. The maintainable CO2 conversions after 160 min of time-on-stream were observed for all feed compositions. However, the obtained CH4 and CO2 conversions in this work were higher than those of the CH4 and CO2 conversions reported in the literature, because of relatively low feed composition and the Ni-based catalyst. For instance, Álvarez M et al. [56] examined the combined dry-steam reforming of methane with the feed composition S/C ratio of 0.4 over the Ni/MgO–Al2O3 catalyst. The received CH4 and CO2 conversions during about 20 h testing time were approximately 35% and 25%, respectively.
The effect of the feed composition towards the syngas production is displayed in Figure 6c; it was observed that the greater S/C ratio in the feed provided the higher H2/CO ratio in the syngas product. The H2/CO ratio of the syngas reached 2 for the S/C ratios of 0.53 and 0.67. The increase of H2 content in the syngas product represented more effect of steam reforming in CSCRM. For all S/C ratios, the H2/CO ratios gained from the tests were slightly lower than the thermodynamic equilibrium values (1.9 for S/C ratio of 0.33, 2.4 for S/C ratio of 0.53, and 2.8 for S/C ratio of 0.67) due to the side reactions such as the RWGS and coke forming reactions [23]. Notwithstanding, these experimental H2/CO ratios refer to a practical feedstock for FTs, as reported in Lillebø et al. [57], who studied the FTs with different H2/CO ratios (between 1.04 and 2.56) on the 20wt%Co-0.5wt%Re/Al2O3 catalyst at 210 °C. The results demonstrated that the hydrocarbon C5+ selectivity of at least 85% can be accomplished for the syngas reactant with the H2/CO ratios above 2.1 without a significant short-term deactivation or a loss of selectivity. Therefore, the feed composition with the S/C ratio of 0.53 accompanied with the operating temperature of 600 °C is a preferable condition for the DCP (consideration of the CH4 conversion and practical consumption) to produce the syngas for FTs in the GTL technology.

2.2.3. TGA/DTG Analysis of Spent Catalysts

The quantity and types of deposited carbon with the effect of the feed composition over the spent NiCo/Mg-Al and the spent Ni/5Ce-Al catalysts in DCP were characterized using the TGA/DTG technique, flowing air, and temperature programmed from 100 to 800 °C. The TGA/DTG profiles are presented in Figure 7a,b for spent NiCo/Mg-Al and Ni/5Ce-Al catalysts, respectively. As depicted in Figure 7a,b, three types of the carbon species were detected. The peaks at the temperature range of 300 to 420 °C are attributed to the oxidation of the weakly stable amorphous carbon (sp2 C-atoms or graphene-like species). The coexistence in two types of carbon species at a high temperature corresponds to the oxidation of carbon nanotubes (450 to 550 °C) and the oxidation of graphitic carbon (sp3 C-atoms, >550 °C) [23]. As seen in Figure 7a, the weight loss of the spent NiCo/Mg-Al catalyst decreased from 65% to 27% with an increasing S/C ratio from 0.33 to 0.53 and then increased from 27% to 33% with an increasing S/C ratio from 0.53 to 0.67. These combined results indicate that the S/C ratio of 0.53 has a tendency to remove a deposited carbon in the case of a low S/C ratio, which can reduce the graphitic carbon. The formation of graphitic carbon existed when adjusting the S/C ratio to equal 0.67. This evidence reflected the longer time of deposited carbon grown on the surface of the NiCo/Mg-Al catalyst for the S/C ratio of 0.67 compared to the S/C ratio of 0.53. It means that the dissociative adsorption of H2O on the NiCo/Mg-Al catalyst requires time before the gasification step to contribute oxygen species in order to remove carbon on the surface. Therefore, fewer active sites for the oxidizing agents were available for the S/C ratio of 0.67 than for the S/C ratio of 0.53, resulting into less oxygen species on the surface, allowing more deposited carbon to polymerization on the surface of Ni.
At the same time, the Ni/5Ce-Al catalyst (Figure 7b) displays the main peak range of 450 to 650 °C, which refers to carbon nanotubes and graphitic carbon. A decrease in weight loss is evident from 25% to 12% with an increasing S/C ratio noted from 0.33 to 0.67. The weight loss for the S/C ratio of above 0.53 is acceptable. From these results, it was found that the content of H2O in the feed composition plays a major role in the removal of carbon species on the surface of the WGS catalyst. By comparison, the UHT-WGS catalyst shows a lower amount of carbon deposition than that of the CSCRM catalyst since the carbon deposition on the Ni-based catalysts is not critical in the WGS reaction [58].

3. Material and Methods

3.1. Catalyst Preparation

3.1.1. CRM Catalyst Preparation

The MgO–Al2O3 support was synthesized by the sol–gel method. Magnesium ethoxide (Mg(OC2H5)2, Sigma Aldrich, St. Louis, MO, USA), was dissolved in distilled water (denoted as solution A) at ambient temperature (25 ° to 30 °C) and stirred for 20 h. At the same time, aluminum isopropoxide (Al[OCH(CH3)2]3, Sigma Aldrich) was dissolved in distilled water (denoted as solution B) and continually stirred at 80 °C for 1 h. Subsequently, concentrated nitric acid was added by drops into the solution B and vigorously stirred for 20 h. Next, solution A was added to solution B and stirred for 2 h to form a homogeneous solution. The obtained mixed solution was dried at 50 °C overnight and calcined at 650 °C for 5 h under an air atmosphere with a ramp rate of 3 °C min−1 to decompose contaminants. The 5wt %Ni5wt %Co/MgO–Al2O3 (NiCo/Mg-Al) catalyst was prepared by co-impregnation using nickel (II) nitrate hexahydrate (Ni(NO3)2·6H2O, Sigma Aldrich) and cobalt (II) nitrate hexahydrate (Co(NO3)2·6H2O, Sigma Aldrich) as precursors for the nickel solution and cobalt solution, respectively. These aqueous solutions were mixed and added by drops onto the support. The obtained solid cake was then dried at 50 °C for 2 h and calcined at 650 °C for 5 h using the ramping rate of 3 °C min−1.

3.1.2. UHT-WGS Catalyst Preparation

The 5wt % CeO2-Al2O3 (Ce-Al) support was first synthesized by the sol–gel method using cerium (III) acetylacetonate hydrate (Ce(C5H7O2)3xH2O, Sigma Aldrich) and aluminum isopropoxide precursors. The desired quantities of the precursors were dissolved in a mixture of distilled water and isopropanol according to a molar ratio of 1:1. The saddle brown solution was refluxed at 70 °C for 2 h to form the gel product. The product was dried at 50 °C overnight and calcined at 600 °C for 6 h with a ramp of 3 °C min−1. Secondly, an amount of nickel (II) nitrate hexahydrate corresponding to 10 wt % of Ni was impregnated onto the calcined support followed by drying at 50 °C for 2 h and calcination at 600 °C for 6 h using the heating rate of 3 °C min−1. For this purpose, the synthesized UHT-WGS catalyst was labeled as Ni/5Ce-Al.

3.2. Catalyst Characterization

3.2.1. Morphological Characterization

The crystalline phases of the catalyst samples were examined by XRD analysis using an X-ray diffractometer (PANalytical X’Pert-Pro, Almelo, The Netherlands) with nickel-filtered Cu K α ( λ   = 1.54178   Å , 2 θ range from 10° to 80°), a monochromatized radiation source, operated at 40 kV and 30 mA, having the scanning rate of 0.02° with 0.5 s per step.
The specific surface area (SBET, m2 g−1), pore volume (Vp, cm3 g−1), and average pore size diameter (nm) were characterized by N2 adsorption/desorption isotherms, which were measured at −196 °C using BELSORP-mini II instrument (Osaka, Japan). The pore size distribution curve was calculated from the analysis of the desorption branch of the isotherm by the BJH method.

3.2.2. Physical Characterization

According to the stoichiometry of 1:1 for the chemisorbed hydrogen atom on the Ni surface, the metal dispersion ( D m , %) and the metal particle size ( d , nm) were calculated from Equations (5) and (6), respectively, using the H2-TPD results [59,60]. In these equations, V chem is an amount of hydrogen desorption (cm3); SF is a stoichiometry factor; MW is an atomic weight of metal (g mol−1); m is a sample weight (g); w is a wt % of supported metal content; σ m is a cross-sectional area of one metal atom (nm2); and ρ   is a density of metal (g cm-3).
% D m = ( V chem 22414 ) × SF × MW ( m × w 100 ) × 100
d = 6000 ( V chem 22414 ) × SF × 6.02 × 10 23 × ρ × σ m × 10 18 ( m × w 100 )
Before analysis, 50 mg of the calcined catalyst was reduced in situ at 600 °C in a H2 flow of 50 mL min−1 for 2 h, followed by cooling to 100 °C in Ar flow of 50 mL min−1. Consequently, H2 was isothermally chemisorbed on the surface of the sample at 100 °C for 0.5 h and the sample was cooled to ambient temperatures in Ar flow of 50 mL min−1. The desorbed H2 was measured by a TCD during the temperature programmed from 40 °C to 900 °C under Ar flow of 50 mL min−1.
The reducibility of the calcined catalyst was evaluated via the H2-TPR technique performed in the BELCAT-basic system (Osaka, Japan). In this analysis, 50 mg of the calcined catalyst was degassed at 220 °C for 1 h in Ar flow of 30 mL min−1, followed by cooling to 40 °C. After, the sample was reduced in the temperature programmed from 40 °C to 900 °C under 5%H2/Ar flow of 50 mL min−1. The H2 consumption was detected by the TCD.
The quantity and nature of the deposited carbon over the spent catalyst were measured by TGA and DTG using a METTLER TOLEDO thermogravimetric analyzer (Columbus, Ohio, United States). The catalyst sample weight loss and the derivative thermogravimetric curve of the weight loss versus temperature were collected continuously under flowing air up to 800 °C with a heating rate of 10 °C min−1.

3.3. Catalytic Activity Test

Catalytic tests were carried out in a stainless steel tubular fixed-bed reactor at atmospheric pressure. Before the DCP reaction, the catalytic performances of CRM and UHT-WGS catalysts were demonstrated separately with a continued reaction temperature programmed (500, 550, and 600 °C); each temperature was held for 8 h. Prior to the CRM test with the composition of CH4:CO2:N2 = 1:1.7:1.3 molar ratio with GHSV of 1.8 × 104 mL gcat−1 h−1, the 200 mg of CRM catalyst diluted with 1000 mg of fused silica was packed (diluted catalyst height of 1.4 cm) and the in situ reduced was at 650 °C for 6 h under a H2 flow of 30 mL min−1; the temperature was then decreased to the reaction temperature in N2 at a flow rate of 30 mL min−1. For the UHT-WGS tests using the GHSV of 2.0 × 105 mL gcat−1 h−1 with H2O/CO ratio of three, the 30 mg of UHT-WGS catalyst diluted with 1000 mg of fused silica (diluted catalyst height of 1.2 cm) was preactivated and cooled using a similar reduction condition to the CRM catalyst.
For the DCP reaction, the diluted UHT-WGS catalyst as previously mentioned was charged first and the quartz wool was then placed on the top of the UHT-WGS catalyst. Subsequently, the diluted CRM catalyst as mentioned was loaded second. When the effect of the operating temperature was evaluated, the different reaction temperatures (500, 550, and 600 °C) were used under the feed composition with the S/C ratio of 0.67. The effect of H2O content in the feed composition was investigated using the feed composition of CH4:CO2:H2O:N2 molar ratio = 1:0.5:x:1; x = 0.5, 0.8, and 1 (corresponding to the S/C ratio of 0.33, 0.53, and 0.67, respectively) with the fixed reaction temperature of 600 °C employing the GHSV rang of 1.6 × 104–1.8 × 104 mL gcat−1 h−1. The conversions of CH4 and CO2 and the H2/CO ratio were calculated using the following Equations (Equations (7)–(9)). The scheme of the experimental setup is presented in Figure 8.
X CH 4 = n ˙ CH 4 , in n ˙ CH 4 , out n ˙ CH 4 , in × 100
X CO 2 = n ˙ CO 2 , in n ˙ CO 2 , out n ˙ CO 2 , in × 100
H 2 CO   ratio = mole   of   H 2   procuced mole   of   CO   procuced

4. Conclusions

The Ni/5Ce-Al catalyst was developed for UHT-WGS and allowed a DCP, including a CSCRM reaction followed by the UHT-WGS reaction over Ni-based catalysts in a single reactor to be created and investigated. For the individual catalyst, the catalytic performances of the CRM and the UHT-WGS catalysts with temperatures programmed from 500 to 600 °C were separately tested. The results revealed that the CH4 conversion in the CRM reaction increased with increasing reaction temperature from 500 to 600 °C because the CRM reaction is an endothermic process that is favored at a high temperature. By contrast, a decrease in the CO conversion in the UHT-WGS reaction with an increasing process temperature from 500 to 600 °C was found due to the exothermic reaction by nature. The convergences of the conversions in the CRM and the UHT-WGS reactions were accomplished at 600 °C. The DCP with the influences of operating temperature (500, 550, and 600 °C) and the feed composition (S/C ratio = 0.33, 0.53, and 0.67) was examined. It was discovered that the operating temperature and the feed composition have significant impacts on conversions, the H2/CO ratio, and the carbon formation in the DCP. The achievement of the optimum DCP condition was a S/C ratio of 0.53 at 600 °C with not only the appropriate H2/CO ratio of about 2 but also the prevention of carbon formation. Therefore, it can be concluded that, at the relative low operating temperature, the DCP on the NiCo/Mg-Al and the Ni/5Ce-Al catalyst could be developed for the commercial production of syngas to be fed to FTs.

Author Contributions

Conceptualization, S.S., T.R., S.T., and M.P.; data curation, S.S. and M.P.; investigation, S.S., T.R., S.T., and M.P.; methodology, S.S. and M.P.; formal analysis, S.S. and M.P.; funding support, T.S.; Writing-original draft, S.S. and M.P.; writing—review and editing, S.S. and M.P.; supervision, M.P., S.T., and E.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Thailand Science Research and Innovation (TSRI) via Research and Researchers for Industries (RRI) with PTT Public Company Limited (grant number PHD59I0028).

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

List of acronyms
BETBrunauer–Emmett–Teller
BJHBarrett–Joyner–Halenda
CRMCO2 reforming of methane
CSCRMcombined steam and CO2 reforming of methane
DCPdual Ni-based catalytic process
DTGderivative thermogravimetric analysis
EISAevaporation-induced self-assembly
FTsFischer–Tropsch synthesis
GHSVgas hourly space velocity
GTLgas-to-liquids
H2-TPDhydrogen temperature programmed desorption
H2-TPRhydrogen temperature programmed reduction
RWGSreverse water gas shift
S/C ratiosteam-to-carbon (H2O/(CH4 + CO2) ratio
SMSIstrong metal support interaction
TCDthermal conductivity detector
TGAthermogravimetric analysis
UHT-WGSultra-high-temperature water–gas shift
XRDX-ray diffraction

References

  1. Koo, K.Y.; Roh, H.S.; Jung, U.H.; Yoon, W.L. Combined H2O and CO2 reforming of CH4over Ce-promoted Ni/Al2O3 catalyst for gas to liquid (GTL) process: Enhancement of Ni-CeO2interaction. Catal. Today 2012, 185, 126–130. [Google Scholar] [CrossRef]
  2. Hadian, N.; Rezaei, M. Combination of dry reforming and partial oxidation of methane over Ni catalysts supported on nanocrystalline MgAl2O4. Fuel 2013, 113, 571–579. [Google Scholar] [CrossRef]
  3. Yang, E.-h.; Noh, Y.S.; Hong, G.H.; Moon, D.J. Combined steam and CO2 reforming of methane over La1-xSrxNiO3 perovskite oxides. Catal. Today 2018, 299, 242–250. [Google Scholar] [CrossRef]
  4. Baek, S.-C.; Bae, J.-W.; Cheon, J.Y.; Jun, K.-W.; Lee, K.-Y. Combined Steam and Carbon Dioxide Reforming of Methane on Ni/MgAl2O4: Effect of CeO2 Promoter to Catalytic Performance. Catal. Lett. 2011, 141, 224–234. [Google Scholar] [CrossRef]
  5. Kumar, R.; Kumar, K.; Choudary, N.V.; Pant, K.K. Effect of support materials on the performance of Ni-based catalysts in tri-reforming of methane. Fuel Process. Technol. 2019, 186, 40–52. [Google Scholar] [CrossRef]
  6. Koo, K.Y.; Roh, H.S.; Jung, U.H.; Seo, D.J.; Seo, Y.S.; Yoon, W.L. Combined H2O and CO2 reforming of CH4 over nano-sized Ni/MgO-Al2O3 catalysts for synthesis gas production for gas to liquid (GTL): Effect of Mg/Al mixed ratio on coke formation. Catal. Today 2009, 146, 166–171. [Google Scholar] [CrossRef]
  7. Roh, H.-S.; Koo, K.Y.; Joshi, U.D.; Yoon, W.L. Combined H2O and CO2 Reforming of Methane Over Ni–Ce–ZrO2 Catalysts for Gas to Liquids (GTL). Catal. Lett. 2008, 125, 283–288. [Google Scholar] [CrossRef]
  8. Koh, A.C.W.; Chen, L.; Kee Leong, W.; Johnson, B.F.G.; Khimyak, T.; Lin, J. Hydrogen or synthesis gas production via the partial oxidation of methane over supported nickel–cobalt catalysts. Int. J. Hydrogen Energy 2007, 32, 725–730. [Google Scholar] [CrossRef]
  9. Rabenstein, G.; Hacker, V. Hydrogen for fuel cells from ethanol by steam-reforming, partial-oxidation and combined auto-thermal reforming: A thermodynamic analysis. J. Power Sources 2008, 185, 1293–1304. [Google Scholar] [CrossRef]
  10. Cavallaro, S.; Chiodo, V.; Vita, A.; Freni, S. Hydrogen production by auto-thermal reforming of ethanol on Rh/Al2O3 catalyst. J. Power Sources 2003, 123, 10–16. [Google Scholar] [CrossRef]
  11. Ma, Q.; Guo, L.; Fang, Y.; Li, H.; Zhang, J.; Zhao, T.S.; Yang, G.; Yoneyama, Y.; Tsubaki, N. Combined methane dry reforming and methane partial oxidization for syngas production over high dispersion Ni based mesoporous catalyst. Fuel Process. Technol. 2019, 188, 98–104. [Google Scholar] [CrossRef]
  12. Nematollahi, B.; Rezaei, M.; Khajenoori, M. Combined dry reforming and partial oxidation of methane to synthesis gas on noble metal catalysts. Int. J. Hydrogen Energy 2011, 36, 2969–2978. [Google Scholar] [CrossRef]
  13. Kim, Y.-H.; Koo, K.-Y.; Song, I.-K. A Simulation Study on SCR (Steam Carbon Dioxide Reforming) Process Optimization for Fischer-Tropsch Synthesis. Korean Chem. Eng. Res. 2009, 47, 700–704. [Google Scholar]
  14. Al-Nakoua, M.A.; El-Naas, M.H. Combined steam and dry reforming of methane in narrow channel reactors. Int. J. Hydrogen Energy 2012, 37, 7538–7544. [Google Scholar] [CrossRef]
  15. Wilhelm, D.; Simbeck, D.; Karp, A.; Dickenson, R. Syngas production for gas-to-liquids applications: Technologies, issues and outlook. Fuel Process. Technol. 2001, 71, 139–148. [Google Scholar] [CrossRef]
  16. Sehested, J. Four challenges for nickel steam-reforming catalysts. Catal. Today 2006, 111, 103–110. [Google Scholar] [CrossRef]
  17. Kumar, N.; Shojaee, M.; Spivey, J. Catalytic bi-reforming of methane: From greenhouse gases to syngas. Curr. Opin. Chem. Eng. 2015, 9, 8–15. [Google Scholar] [CrossRef] [Green Version]
  18. Olah, G.A.; Goeppert, A.; Czaun, M.; Mathew, T.; May, R.B.; Prakash, G.K.S. Single Step Bi-reforming and Oxidative Bi-reforming of Methane (Natural Gas) with Steam and Carbon Dioxide to Metgas (CO-2H2) for Methanol Synthesis: Self-Sufficient Effective and Exclusive Oxygenation of Methane to Methanol with Oxygen. J. Am. Chem. Soc. 2015, 137, 8720–8729. [Google Scholar] [CrossRef]
  19. Olah, G.A.; Goeppert, A.; Czaun, M.; Prakash, G.K.S. Bi-reforming of Methane from Any Source with Steam and Carbon Dioxide Exclusively to Metgas (CO–2H2) for Methanol and Hydrocarbon Synthesis. J. Am. Chem. Soc. 2013, 135, 648–650. [Google Scholar] [CrossRef]
  20. Jang, W.-J.; Jeong, D.-W.; Shim, J.-O.; Kim, H.-M.; Roh, H.-S.; Son, I.H.; Lee, S.J. Combined steam and carbon dioxide reforming of methane and side reactions: Thermodynamic equilibrium analysis and experimental application. Appl. Energy 2016, 173, 80–91. [Google Scholar] [CrossRef]
  21. Danilova, M.M.; Fedorova, Z.A.; Kuzmin, V.A.; Zaikovskii, V.I.; Porsin, A.V.; Krieger, T.A. Combined steam and carbon dioxide reforming of methane over porous nickel based catalysts. Catal. Sci. Technol. 2015, 5, 2761–2768. [Google Scholar] [CrossRef]
  22. Huang, B.; Li, X.; Ji, S.; Lang, B.; Habimana, F.; Li, C. Effect of MgO promoter on Ni-based SBA-15 catalysts for combined steam and carbon dioxide reforming of methane. J. Nat. Gas Chem. 2008, 17, 225–231. [Google Scholar] [CrossRef]
  23. Jabbour, K.; Massiani, P.; Davidson, A.; Casale, S.; El Hassan, N. Ordered mesoporous “one-pot” synthesized Ni-Mg(Ca)-Al2O3 as effective and remarkably stable catalysts for combined steam and dry reforming of methane (CSDRM). Appl. Catal. B Environ. 2017, 201, 527–542. [Google Scholar] [CrossRef] [Green Version]
  24. Bae, J.W.; Kim, A.R.; Baek, S.-C.; Jun, K.-W. The role of CeO2–ZrO2 distribution on the Ni/MgAl2O4 catalyst during the combined steam and CO2 reforming of methane. React. Kinet. Mech. Catal. 2011, 104, 377–388. [Google Scholar] [CrossRef]
  25. Yagi, F.; Kanai, R.; Wakamatsu, S.; Kajiyama, R.; Suehiro, Y.; Shimura, M. Development of synthesis gas production catalyst and process. Catal. Today 2005, 104, 2–6. [Google Scholar] [CrossRef]
  26. Samavati, M.; Santarelli, M.; Martin, A.; Nemanova, V. Thermodynamic and economy analysis of solid oxide electrolyser system for syngas production. Energy 2017, 122, 37–49. [Google Scholar] [CrossRef]
  27. Selvatico, D.; Lanzini, A.; Santarelli, M. Low Temperature Fischer-Tropsch fuels from syngas: Kinetic modeling and process simulation of different plant configurations. Fuel 2016, 186, 544–560. [Google Scholar] [CrossRef]
  28. Iaquaniello, G.; Salladini, A.; Palo, E.; Centi, G. Catalytic Partial Oxidation Coupled with Membrane Purification to Improve Resource and Energy Efficiency in Syngas Production. ChemSusChem 2015, 8, 717–725. [Google Scholar] [CrossRef] [PubMed]
  29. Asencios, Y.J.O.; Assaf, E.M. Combination of dry reforming and partial oxidation of methane on NiO-MgO-ZrO2 catalyst: Effect of nickel content. Fuel Process. Technol. 2013, 106, 247–252. [Google Scholar] [CrossRef]
  30. Li, X.; Teng, Y.; Gong, M.; Chen, Y. Combination of partial oxidation and CO2 reforming of methane over monolithic Ni/CeO2–ZrO2/γ-Al2O3 Catalyst. Chin. Chem. Lett. 2005, 16, 691–692. [Google Scholar]
  31. Piyapaka, K.; Tungkamani, S.; Phongaksorn, M. Effect of Strong Metal Support Interactions of Supported Ni and Ni-Co Catalyst on Metal Dispersion and Catalytic Activity toward Dry Methane Reforming Reaction. King Mongkut’s Univ. Technol. North Bangkok Int. J. Appl. Sci. Technol. 2016, 9, 255–259. [Google Scholar] [CrossRef] [Green Version]
  32. Sangsong, S.; Ratana, T.; Tungkamani, S.; Sornchamni, T.; Phongaksorn, M. Effect of CeO2 loading of the Ce-Al mixed oxide on ultrahigh temperature water-gas shift performance over Ce-Al mixed oxide supported Ni catalysts. Fuel 2019, 252, 488–495. [Google Scholar] [CrossRef]
  33. Chen, L.; Zhu, Q.; Wu, R. Effect of Co–Ni ratio on the activity and stability of Co–Ni bimetallic aerogel catalyst for methane Oxy-CO2 reforming. Int. J. Hydrogen Energy 2011, 36, 2128–2136. [Google Scholar] [CrossRef]
  34. Akbari, E.; Alavi, S.M.; Rezaei, M. Synthesis gas production over highly active and stable nanostructured NiMgOAl2O3 catalysts in dry reforming of methane: Effects of Ni contents. Fuel 2017, 194, 171–179. [Google Scholar] [CrossRef]
  35. Katheria, S.; Gupta, A.; Deo, G.; Kunzru, D. Effect of calcination temperature on stability and activity of Ni/MgAl2O4 catalyst for steam reforming of methane at high pressure condition. Int. J. Hydrogen Energy 2016, 41, 14123–14132. [Google Scholar] [CrossRef]
  36. Profeti, L.P.R.; Ticianelli, E.A.; Assaf, E.M. Production of hydrogen via steam reforming of biofuels on Ni/CeO2–Al2O3 catalysts promoted by noble metals. Int. J. Hydrogen Energy 2009, 34, 5049–5060. [Google Scholar] [CrossRef]
  37. Rojas, H.; Borda, G.; Reyes, P.; Brijaldo, M.; Valencia, J. Liquid-phase hydrogenation of m-dinitrobenzene over platinum catalysts. J. Chil. Chem. Soc. 2011, 56, 793–798. [Google Scholar] [CrossRef] [Green Version]
  38. Ferreira, A.P.; Zanchet, D.; Rinaldi, R.; Schuchardt, U.; Damyanova, S.; Bueno, J.M.C. Effect of the CeO2 content on the surface and structural properties of CeO2-Al2O3 mixed oxides prepared by sol-gel method. Appl. Catal. A Gen. 2010, 388, 45–56. [Google Scholar] [CrossRef]
  39. Horlyck, J.; Lawrey, C.; Lovell, E.C.; Amal, R.; Scott, J. Elucidating the impact of Ni and Co loading on the selectivity of bimetallic NiCo catalysts for dry reforming of methane. Chem. Eng. J. 2018, 352, 572–580. [Google Scholar] [CrossRef]
  40. Iriondo, A.; Barrio, V.L.; Cambra, J.F.; Arias, P.L.; Guemez, M.B.; Sanchez-Sanchez, M.C.; Navarro, R.M.; Fierro, J.L.G. Glycerol steam reforming over Ni catalysts supported on ceria and ceria-promoted alumina. Int. J. Hydrogen Energy 2010, 35, 11622–11633. [Google Scholar] [CrossRef]
  41. Luisetto, I.; Tuti, S.; Battocchio, C.; Lo Mastro, S.; Sodo, A. Ni/CeO2–Al2O3 catalysts for the dry reforming of methane: The effect of CeAlO3 content and nickel crystallite size on catalytic activity and coke resistance. Appl. Catal. A Gen. 2015, 500, 12–22. [Google Scholar] [CrossRef]
  42. Li, Z.; Li, M.; Ashok, J.; Kawi, S. NiCo@NiCo phyllosilicate@CeO2 hollow core shell catalysts for steam reforming of toluene as biomass tar model compound. Energy Convers. Manag. 2019, 180, 822–830. [Google Scholar] [CrossRef]
  43. Sirijaruphan, A.; Horváth, A.; Goodwin Jr., J.G.; Oukaci, R. Cobalt Aluminate Formation in Alumina-Supported Cobalt Catalysts: Effects of Cobalt Reduction State and Water Vapor. Catal. Lett. 2003, 91, 89–94. [Google Scholar] [CrossRef]
  44. Patcas, F.; Hönicke, D. Effect of alkali doping on catalytic properties of alumina-supported nickel oxide in the selective oxidehydrogenation of cyclohexane. Catal. Commun. 2005, 6, 23–27. [Google Scholar] [CrossRef]
  45. Ashok, J.; Kawi, S. Steam reforming of toluene as a biomass tar model compound over CeO2 promoted Ni/CaO-Al2O3 catalytic systems. Int. J. Hydrogen Energy 2013, 38, 13938–13949. [Google Scholar] [CrossRef]
  46. Djinović, P.; Osojnik Črnivec, I.G.; Erjavec, B.; Pintar, A. Influence of active metal loading and oxygen mobility on coke-free dry reforming of Ni–Co bimetallic catalysts. Appl. Catal. B Environ. 2012, 125, 259–270. [Google Scholar] [CrossRef]
  47. Tanios, C.; Bsaibes, S.; Gennequin, C.; Labaki, M.; Cazier, F.; Billet, S.; Tidahy, H.L.; Nsouli, B.; Aboukaïs, A.; Abi-Aad, E. Syngas production by the CO2 reforming of CH4 over Ni–Co–Mg–Al catalysts obtained from hydrotalcite precursors. Int. J. Hydrogen Energy 2017, 42, 12818–12828. [Google Scholar] [CrossRef]
  48. Jha, A.; Jeong, D.-W.; Jang, W.-J.; Rode, C.V.; Roh, H.-S. Mesoporous NiCu–CeO2 oxide catalysts for high-temperature water–gas shift reaction. RSC Adv. 2015, 5, 1430–1437. [Google Scholar] [CrossRef]
  49. Haryanto, A.; Fernando, S.; Adhikari, S. Ultrahigh temperature water gas shift catalysts to increase hydrogen yield from biomass gasification. Catal. Today 2007, 129, 269–274. [Google Scholar] [CrossRef]
  50. Oemar, U.; Bian, Z.; Hidajat, K.; Kawi, S. Sulfur resistant LaxCe1−xNi0.5Cu0.5O3 catalysts for an ultra-high temperature water gas shift reaction. Catal. Sci. Technol. 2016, 6, 6569–6580. [Google Scholar] [CrossRef]
  51. Li, L.; Wang, H.; Jiang, X.; Song, Z.; Zhao, X.; Ma, C. Microwave-enhanced methane combined reforming by CO2 and H2O into syngas production on biomass-derived char. Fuel 2016, 185, 692–700. [Google Scholar] [CrossRef]
  52. Gensterblum, Y.; Busch, A.; Krooss, B.M. Molecular concept and experimental evidence of competitive adsorption of H2O, CO2 and CH4 on organic material. Fuel 2014, 115, 581–588. [Google Scholar] [CrossRef]
  53. Tan, K.; Zuluaga, S.; Gong, Q.; Gao, Y.; Nijem, N.; Li, J.; Thonhauser, T.; Chabal, Y.J. Competitive coadsorption of CO2 with H2O, NH3, SO2, NO, NO2, N2, O2, and CH4 in M-MOF-74 (M = Mg, Co, Ni): The role of hydrogen bonding. Chem. Mater. 2015, 27, 2203–2217. [Google Scholar] [CrossRef] [Green Version]
  54. Zhou, W.; Wang, H.; Zhang, Z.; Chen, H.; Liu, X. Molecular simulation of CO2/CH4/H2O competitive adsorption and diffusion in brown coal. RSC Adv. 2019, 9, 3004–3011. [Google Scholar] [CrossRef] [Green Version]
  55. Singh, S.; Bahari, M.B.; Abdullah, B.; Phuong, P.T.T.; Truong, Q.D.; Vo, D.V.N.; Adesina, A.A. Bi-reforming of methane on Ni/SBA-15 catalyst for syngas production: Influence of feed composition. Int. J. Hydrogen Energy 2018, 43, 17230–17243. [Google Scholar] [CrossRef]
  56. Álvarez M, A.; Centeno, M.Á.; Odriozola, J.A. Ru-Ni Catalyst in the Combined Dry-Steam Reforming of Methane: The Importance in the Metal Order Addition. Top. Catal. 2016, 59, 303–313. [Google Scholar] [CrossRef]
  57. Lillebø, A.; Rytter, E.; Blekkan, E.A.; Holmen, A. Fischer-Tropsch Synthesis at High Conversions on Al2O3-Supported Co Catalysts with Different H2/CO Levels. Ind. Eng. Chem. Res. 2017, 56, 13281–13286. [Google Scholar] [CrossRef]
  58. Ashok, J.; Wai, M.H.; Kawi, S. Nickel-based Catalysts for High-temperature Water Gas Shift Reaction-Methane Suppression. ChemCatChem 2018, 10, 3927–3942. [Google Scholar] [CrossRef]
  59. Prakash, N.; Lee, M.H.; Yoon, S.; Jung, K.D. Role of acid solvent to prepare highly active PtSn/Θ-Al2O3 catalysts in dehydrogenation of propane to propylene. Catal. Today 2017, 293–294, 33–41. [Google Scholar] [CrossRef]
  60. Barbier, A.; Tuel, A.; Martin, G.A.; Arcon, I.; Arcon, I.; Kodre, A.; Kodre, A. Characterization and catalytic behavior of CO/SiO2 catalysts: Influence of dispersion in the Fischer-Tropsch reaction. J. Catal. 2001, 200, 106–116. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD patterns of calcined catalysts.
Figure 1. XRD patterns of calcined catalysts.
Catalysts 10 01056 g001
Figure 2. N2 adsorption–desorption isotherms and BJH pore distributions of calcined. (a) NiCo/Mg-Al and (b) Ni/5Ce-Al catalysts.
Figure 2. N2 adsorption–desorption isotherms and BJH pore distributions of calcined. (a) NiCo/Mg-Al and (b) Ni/5Ce-Al catalysts.
Catalysts 10 01056 g002
Figure 3. H2-TPR profiles of calcined catalysts.
Figure 3. H2-TPR profiles of calcined catalysts.
Catalysts 10 01056 g003
Figure 4. (a) CH4 conversion over the NiCo/Mg-Al catalyst in the CRM and (b) CO conversion over the Ni/5Ce-Al catalyst in the UHT-WGS with a continue reaction temperature programmed. (500 °C to 600 °C).
Figure 4. (a) CH4 conversion over the NiCo/Mg-Al catalyst in the CRM and (b) CO conversion over the Ni/5Ce-Al catalyst in the UHT-WGS with a continue reaction temperature programmed. (500 °C to 600 °C).
Catalysts 10 01056 g004
Figure 5. (a) CH4 conversion, (b) CO2 conversion, and (c) H2/CO ratio with the effect of operating temperature for the DCP with a fixed S/C ratio of 0.67.
Figure 5. (a) CH4 conversion, (b) CO2 conversion, and (c) H2/CO ratio with the effect of operating temperature for the DCP with a fixed S/C ratio of 0.67.
Catalysts 10 01056 g005
Figure 6. The effect of the feed composition (S/C ratio = 0.33, 0.53, and 0.67) in the DCP at a fixed temperature of 600 °C; (a) CH4 conversion, (b) CO2 conversion, and (c) H2/CO ratio.
Figure 6. The effect of the feed composition (S/C ratio = 0.33, 0.53, and 0.67) in the DCP at a fixed temperature of 600 °C; (a) CH4 conversion, (b) CO2 conversion, and (c) H2/CO ratio.
Catalysts 10 01056 g006
Figure 7. TGA/DTG profiles of spent (a) NiCo/Mg-Al and (b) Ni/5Ce-Al catalysts for the effect of feed composition at 600 °C.
Figure 7. TGA/DTG profiles of spent (a) NiCo/Mg-Al and (b) Ni/5Ce-Al catalysts for the effect of feed composition at 600 °C.
Catalysts 10 01056 g007
Figure 8. Schematic diagram of the experimental setup.
Figure 8. Schematic diagram of the experimental setup.
Catalysts 10 01056 g008
Table 1. Physicochemical properties of calcined catalysts.
Table 1. Physicochemical properties of calcined catalysts.
CatalystsSBET
(m2 g−1) a
Vp (cm3 g−1) a Average Pore Size Diameter (nm) a% D m   b   d   ( nm )   b H2-Uptakes (µmol g−1)
Actual cTheoretical d
NiCo/Mg-Al1300.411.232.22.016161700
Ni/5Ce-Al1830.612.523.82.714281704
a Calculated by the BET Equation with about 5% systematic error; b calculated from H2-TPD results with about 8% systematic error; H2 consumption calculated experimentally (Actual c) from TPR profiles (with about 8% systematic error) after complete reduction at T = 900 °C and theoretical values (Theoretical d) determined based on metal loading.

Share and Cite

MDPI and ACS Style

Sangsong, S.; Ratana, T.; Tungkamani, S.; Sornchamni, T.; Phongaksorn, M.; Croiset, E. The Demonstration of the Superiority of the Dual Ni-Based Catalytic System for the Adjustment of the H2/CO Ratio in Syngas for Green Fuel Technologies. Catalysts 2020, 10, 1056. https://doi.org/10.3390/catal10091056

AMA Style

Sangsong S, Ratana T, Tungkamani S, Sornchamni T, Phongaksorn M, Croiset E. The Demonstration of the Superiority of the Dual Ni-Based Catalytic System for the Adjustment of the H2/CO Ratio in Syngas for Green Fuel Technologies. Catalysts. 2020; 10(9):1056. https://doi.org/10.3390/catal10091056

Chicago/Turabian Style

Sangsong, Suntorn, Tanakorn Ratana, Sabaithip Tungkamani, Thana Sornchamni, Monrudee Phongaksorn, and Eric Croiset. 2020. "The Demonstration of the Superiority of the Dual Ni-Based Catalytic System for the Adjustment of the H2/CO Ratio in Syngas for Green Fuel Technologies" Catalysts 10, no. 9: 1056. https://doi.org/10.3390/catal10091056

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop